Download (Neoproterozoic-Paleozoic) and plate reconstructions

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Yilgarn Craton wikipedia , lookup

Oceanic trench wikipedia , lookup

Great Lakes tectonic zone wikipedia , lookup

Plate tectonics wikipedia , lookup

Algoman orogeny wikipedia , lookup

Large igneous province wikipedia , lookup

Baltic Shield wikipedia , lookup

Supercontinent wikipedia , lookup

Pangaea wikipedia , lookup

Cimmeria (continent) wikipedia , lookup

Transcript
Early evolution stages of the arctic margins
(Neoproterozoic-Paleozoic) and plate reconstructions
V. A. Vernikovsky1, 2, D. V. Metelkin1, 2, A. E. Vernikovskaya1, N. Yu. Matushkin1, 2, L. I. Lobkovsky3, E.
V. Shipilov4
Trofimuk Institute of Petroleum Geology and Geophysics, Siberian Branch of the RAS, 3, Akademika Koptyuga
Prosp., Novosibirsk, Russia, 630090
2
Novosibirsk State University, 2, Pirogova St., Novosibirsk, Russia, 630090
3
Shirshov Institute of Oceanology of the RAS, 36, Nahimovsky Prosp., Moscow, Russia, 117997
4
Polar Geophysical Institute, Kola Science Centre of the RAS, 15, Khalturina St., Murmansk, Russia, 183010
1
ABSTRACT
In this paper we offer paleoreconstructions
for key structures of the Arctic based on the
synthesis of geostructural, geochronological and
new paleomagnetic data bearing upon the Late
Neoproterozoic and the Paleozoic histories of the
Taimyr fold belt and Kara microcontinent. These
tectonic features are part of a greater continental mass
that we term “Arctida”, with an interesting history of
breakup and reassembly that is constrained by our new
data and synthesis. In the Central Taimyr accretionary
belt fragments of an ancient island arc (960 Ma)
have been discovered, and the paleomagnetic pole
for the arc approximates the synchronous (950 Ma)
pole for the Siberian paleocontinent. For the Kara
microcontinent we demonstrate its evolution in the
Early Paleozoic and its collision with Siberia in the
Late Paleozoic. These data along with an extensive
published paleomagnetic database for the cratons of
Laurentia, Baltica, Siberia, and Gondwana are the
basis for the presented paleotectonic reconstructions.
The migrations of those Arctida tectonic blocks that
lack paleomagnetic data are reconstructed based on
geologic information.
INTRODUCTION
The current structure of the Arctic Ocean
is determined by the position of the Amerasian
(Canadian) and Eurasian basins, whose formation
took place as a result of significant tectonic processes
in the Late Mesozoic – Cenozoic. However it is
impossible to understand relatively recent and
modern tectonic displacements without analyzing
previous tectonic events.
The discovery of Precambrian metamorphic
complexes among the main structures of the
ICAM VI Proceedings
Arctic Region led to the suggestion that in the
Late Precambrian a paleocontinent – termed
“Arctida” – existed between Laurentia, Baltica and
Siberia (Zonenshain, Natapov, 1987). In the classic
presentation it is composed of several blocks of
continental crust, whose relicts are now located in
the Arctic (Fig. 1): the Kara block, the New Siberian
block (the New Siberian Islands and the adjacent
shelf), the North Alaska and Chukotka blocks,
as well as small fragments of the Inuit Fold Belt
in northern Greenland (Peary Land, the northern
part of Ellesmere and Axel Heiberg islands) and
the blocks of the underwater Lomonosov and
Alpha-Mendeleev Ridges (Zonenshain, Natapov,
1987; Zonenshain et al., 1990). In the modern
interpretation, aside from these fragments, Arctida
also includes parts of Barentsia, which includes the
structures of the Svalbard and the Timan-Pechora
plates (Vernikovsky, 1996; Kuznetsov et al., 2007).
Late Precambrian and Paleozoic global tectonic
history is defined by the breakup of Rodinia, the
evolution of newly formed oceanic basins and the
formation of Pangea as a result. Many paleotectonic
schemes and reconstructions have been composed
for the Late Precambrian – Paleozoic stages of the
plates interactions (Scotese and McKerrow, 1990;
Dalziel, 1991,1997; Hoffman, 1991; Powell et
al., 1993; Condie and Rosen, 1994; Torsvik et al.,
1996; Golonka, 2002; Golonka et al., 2003; Cocks
and Torsvik, 2002; Lawver et al., 2002; Li et al.,
2008; Pisarevsky et al., 2008, Metelkin et al., 2012).
However, when dealing with the details of the
evolution of separate lithosphere segments, including
those of the Arctic Region, there are still many
unsolved, debatable and ill-founded reconstructions.
This is true mainly for the deciphering of the initial
265
Fig. 1. (a) The main blocks, microcontinents, plates, and basins of the Arctic on the International Bathymetric Chart
of the Arctic Ocean and (b) a reconstruction for the Early Jurassic, showing the Precambrian Arctic blocks (in red),
amalgamated into the Arctida continent, which is attached to Laurasia (Zonenshain and Natapov, 1987; Zonenshain et
al., 1990). The approximate location of the field study area within the Taimyr folded area is shown by orange dots.
structure of Arctida, the reasons and mechanisms of
its breakup, the drift trajectories of the continental
blocks that composed it. The very existence of oceanic
basins that supposedly separated the paleocontinents
is uncertain. All these are largely debatable topics,
especially the early stages of the Arctic Region
tectonic evolution – the Late Precambrian and the
Early Paleozoic, which are the subject of this paper.
In this study we have attempted to integrate the
available geologic and geophysical material for the
early evolution stages of the Arctic Ocean in the
form of a series of paleotectonic reconstructions, as
well as to create a new development model for the
structures of Arctida.
The determination of the relative positions of
the blocks composing Arctida could be done with
paleomagnetic data. However, such data are very
sparse for the Late Precambrian and the Paleozoic.
For the entire Arctic Region the IAGA Global
Paleomagnetic Database counts no more than 30
paleomagnetic determinations. Nearly all of the
available data represent the Late Paleozoic and
Early Mesozoic of the Barentsia and GreenlandEllesmere regions. There are no data for the New
Siberian Islands and the territories of Chukotka and
Northern Alaska, which represent most of the classic
266
Arctida area. Reliable paleomagnetic determinations
for the Neoproterozoic-Paleozoic time interval are
available only for fragments of a 960 Ma island arc
from Central Taimyr (Vernikovsky et al., 2011) and
for which the paleomagnetic pole is comparable
to the approximately synchronous pole of Siberia
from (Pavlov et al., 2002). There are other reliable
data for the Kara microcontinent: this includes
three paleomagnetic poles for 500, 450 and 420 Ma
(Metelkin et al., 2000; 2005). It is these data that are
placed at the core of our paleotectonic reconstructions
along with the extensive paleomagnetic database for
the Laurentia, Baltica, Siberia and Gondwana cratons
(Pechersky and Didenko, 1995; Torsvik et al., 1996;
Smethurst et al., 1998; McElhinny and MacFadden,
2000; Wingate and Giddings, 2000; Pavlov et al.,
2002; Torsvik and Van der Voo, 2002; Meert and
Torsvik, 2003; Metelkin et al., 2007, 2012; Li et al.,
2008). The paleogeographic position of the cratons
is corrected (within confidence limits for paleopoles)
in accordance with the general model and available
global reconstructions, including structures of the
Arctic sector (Scotese, 1997; Lawver et al., 2002,
2011; Golonka et al., 2003, 2006; Kurenkov et al.,
2005; Cocks and Torsvik, 2002, 2007).
V. A. Vernikovsky
THE OLDEST ISLAND ARC COMPLEX OF
CENTRAL TAIMYR
The Central-Taimyr accretionary belt is located
between two large continental blocks – the Siberian
craton on the south and the Kara microcontinent on
the north (Fig. 2a). It is composed of paleo-island arc
fragments, granite-metamorphic terranes, passive
continental margin terranes of mainly carbonate
composition and ophiolites, which were amalgamated
and accreted to the Siberian craton in the Late
Neoproterozoic and then unconformably overlain
by a Vendian (Ediacaran) – Early Carboniferous
cover (Uflyand et al., 1991; Vernikovsky, 1996;
Khain et al., 1997; Vernikovsky and Vernikovskaya,
2001; Pease et al., 2001). In this model a significant
role is played by ophiolites and island arcs, whose
zircons U–Pb age has been established in the interval
of 755–730 Ma from plagiogranites, gabbros and
volcanogenic rocks (Vernikovsky et al., 1994; 2004).
However, no paleomagnetic data have been obtained
for the 755–730 Ma rocks. Investigations carried out
in the North-Eastern Taimyr in recent years allowed
us to identify an older (960 Ma) paleo-island arc
complex in the Central Taimyr accretionary belt and
to establish its location at the time of formation by
using paleomagnetic data.
The studied area of the Three Sisters Lake (Fig.
2b) is the junction zone of Zhdanov formation rocks
(Zabiyaka et al., 1986) and mainly volcanogenic
rocks previously included in the Borzov (Bezzubtsev
et al., 1986) or Laptev (Zabiyaka et al., 1986)
formations. Zhdanov formation rocks are mainly
terrigeneous (greenish-grey and grey sandstones,
siltstones, black and dark-grey phyllites with
separate layers of carbonate rocks, andesite-basalts,
acid effusive rocks and their tuffs), metamorphosed
in greenschist facies conditions. Borzov/Laptev
formation rocks (metamorphosed basalts, andesites,
dacites, plagiorhyodacites) are host to plagiogranites
and plagiogranite porphyry. Both formations are
intruded by slightly metamorphosed gabbro-dolerite
sills and dikes with thicknesses ranging from tens of
centimeters to hundreds of meters which compose
a wide dike belt with a total length of over 100 km.
The Zhdanov and Borzov/Laptev formations are
overlain by the coarse-grained terrigeneous deposits
of the Oktyabrsk formation. In the studied region the
rocks that compose the island arc are tectonically
ICAM VI Proceedings
composited with sedimentary and volcanogenicsedimentary deposits that we consider to have
formed in an adjoining back-arc basin.
In the study area andesites, dacites, and
plagiorhyodacites are the dominant rock types in
the paleo-island arc complex. These rocks range
from dark-grey with a lilac hue to bottle-green,
sometimes with 2–4 mm phenocrysts of plagioclase,
quartz and less frequently of subordinate potassium
feldspar. Andesites are distinguished by their finegrained matrix textures and by the presence of suites
of ore minerals. Hypabyssal rocks are represented
by metamorphosed plagiogranite porphyry with
medium-grained matrix texture and consisting
mainly of quartz and plagioclase. Plagiogranites also
contain hornblende and clinopyroxene. Secondary
minerals are albite, chlorite, biotite, carbonate,
and epidote. These rocks are often schistose and
highly fractured and veined. Diabases and gabbrodolerites are fine- and medium-grained and intensely
amphibolized.
The studied acid-intermediate volcanic
and intrusive rocks are attributed to the tholeitic
magmatic series. They have weakly or moderately
fractionated REE spectra ((La/Yb)N = 3.3–11.5) with
small negative Eu anomalies (Eu/Eu* = 0.7–0.9),
the total REE concentration is 330–781 ppm. On
the spider diagrams the rocks are enriched in La,
Ce and also Th and U and depleted in Sr, Ti, P, Ta,
and Nb. For the island arc metabasites the total REE
concentration varies in a wider interval from 233–
375 to 1290 ppm. They can have small Eu anomalies
(Eu/Eu* = 0.9–1.1), whereas the (La/Yb)N ratio
values vary widely from 1–3.5 to 36.6. The REE
spectra are flat, typical of MORB and close to those
of island arc basalts. The established Rare Earth and
other trace elements distribution types for the entire
complex are similar to those of the igneous rocks of
Neoproterozoic island arc of other Taimyr regions
(Vernikovsky et al, 1994, 2004).
We performed U–Pb isotopic analysis utilizing a
multicollector Finnigan MAT-261 mass spectrometer
and the Sm, Nd, Rb, and Sr analysis – on a 7-collector
Triton T1 mass spectrometer at the Institute of
Precambrian Geology and Geochronology of the
RAS, St. Petersburg (Russia).
The accessory zircons from a plagiorhyodacite
and a plagiogranite are semitransparent and
267
Fig. 2. (a) The Three Sisters Lake study area location on the tectonic scheme of the Taimyr folded area; and (b) a
geologic map of the Three Sisters Lake study area composed using the data of Bezzubtsev et al., (1986); Zabiyaka et al.,
(1986); and Vernikovsky (1996). 1–6 – Tectonic elements and geodynamic complexes on the tectonic scheme shown on
the regional map (a): 1 – Kara microcontinent (NP–PZ); 2 – collisional granitoids (300–264 Ma, after Vernikovsky et
al., (1995; 1998)); 3 – Central Taimyr accretionary belt (NP) including 4 – Mamont-Shrenk (1) and Faddey (2) cratonic
terranes; 5 – South Taimyr folded belt (PZ–MZ); 6 – overlapping sedimentary complex. 7–16 – Neoproterozoic rocks
shown on the geological map (b): 7–9 – Zhdanov formation including: 7 – black phyllites and siltstones; 8 – sandstones
and siltstones with subordinate interbeds of phyllites; 9 – lenses of limestones and dolomites; 10–11 – Borzov/Laptev
formation including: 10 – andesites, dacites, subordinate basalts and andesite-basalts; 11 – plagiorhyodacites; 12 –
intrusions of plagiogranites; 13 – gabbro-dolerite sill; 14–16 – overlapping strata of Oktyabrsk formation including:
14 – quartz and polymict conglomerates; 15 – oligomict and quartz sandstones and gritstones; 16 – polymict and quartz
conglomerates, breccias. 17–21 – faults and other symbols shown in both maps (a) and (b): 17 – sutures: I – Main Taimyr,
II – Diabasovy, III – Pyasina-Faddey; 18 – normal faults, reverse faults, strike slip faults, 19 – thrusts; 20 – inferred
faults; 21 – sampling sites for geochronological (red) and paleomagnetic (yellow) investigations; 22 – strata bedding.
268
V. A. Vernikovsky
transparent subidiomorphic pink crystals of prismatic
and short-prismatic shape. The morphological
particularities of the zircon grains indicate their
magmatic origin. The isotopic composition points
of the studied zircons from a plagiorhyodacite
(sample A02-16) are approximated by a regression
line, where the upper intersection with the concordia
corresponds to the age 966±5 Ma and the lower
intersection corresponds to 279±30 Ma, with MSWD
= 0.84 (Vernikovsky et al., 2011). At the same
time the isotopic composition points for the zircon
residue after acid treatment with longer exposition
is located on the concordia, and its concordia age is
961±3 Ma (MSWD = 0.72, probability = 0.4) and
can be accepted as the most precise crystallization
time estimate for the studied zircons.
The isotopic composition points for 20
untreated zircon grains from a plagiogranite (sample
A02-2) and for two residues after acid treatment
form a discordia whose upper intersection with the
concordia corresponds to the age of 989±41 Ma, and
the lower one – 508±410 Ma, MSWD = 0.05. The
mean age value, calculated from the 207Pb/206Pb ratio
of the three fractions of the studied zircon grains
correspond to 969±17 Ma and is close to the age
value obtained from the upper intersection with the
discordia. This age estimate may be used as the most
precise one (Vernikovsky et al., 2011).
Sm–Nd isotopic data for island arc acid intrusive
and volcanic rocks of the Three Sisters Lake region
yield a Mesoproterozoic model age: TNd(DM) varies
from 1170 to 1219 Ma. These data as well as Rb–
Sr isotopic investigations indicate a predominance
of a mantle component in the magmatic sources of
these rocks: εNd(967–961) = 5.1–5.2 and (87Sr/86Sr0) =
0.70258–0.70391 (Vernikovsky et al., 2011).
The paleomagnetic analysis was performed on
the apparatus of the Paleomagnetic Center in the
Laboratory of Geodynamics and Paleomagnetism
of the IPGG SB RAS (Novosibirsk). The hardware
system comprises new generation measurers
including a 2G Enterprises Superconductive
Magnetometer (USA) with built-in AF-demagnetizer
and an HSM superconductive spinner-magnetometer
(Germany), as well as the well-known JR-4 and
JR-6 spin-magnetometers (Czech Republic) and
other instruments, placed in a shielded room.
The investigation includes a detailed stepwise
ICAM VI Proceedings
thermal
demagnetization
(T-demagnetization)
and/or alternating field demagnetization (AFdemagnetization) of all studied samples until their
complete demagnetization.
The experimental results were processed with
specialized software products that use standard
techniques for component analysis (Butler,
1992); and various modifications of the fold test
(McFadden, 1990; Watson and Enkin, 1993, Enkin,
2003) and reversal test (McFadden and McElhinny,
1990) for dating the magnetization components.
The sample collection includes volcanic as well as
intrusive rocks of the paleo-island arc complex (Fig.
2b). One site (02ta-4) corresponds to an outcrop
of plagiorhyodacites (sample A02-16), which has
been dated by U-Pb method. The studied rocks are
characterized by relatively low values of natural
remnant magnetization, NRM (tens of mA/m,
thousands for one outcrop) and a high magnetic
susceptibility - about 10-3 SI units. For the analysis of
the NRM components T-, and AF-demagnetization
were used. Typical orthogonal plots are given in
Fig. 3. Most of the samples are characterized by two
often unidirectional components – a titanomagnetite
component with a blocking temperature TB of
about 400°C and a magnetite component with TB ~
580°C. Distinctive particularities in the NRM vector
behavior during the demagnetization of rocks from
various outcrops are mainly due to the input of the
titanomagnetite and magnetite components. In some
samples the component of characteristic remnant
magnetization (ChRM; shown as dashed lines in
vertical plane projections (open circles), Fig. 3) is
exactly registered in a high temperature interval 400–
580°C, and the almost complete demagnetization
of others is reached with the heating to 400°C or
lower. In the last case (lower right in Fig. 3) the
AF-demagnetization is more informative. The value
of the median destructive field (MDF) is no more
than 20–30 mT, and the complete demagnetization
is reached by the impact of the alternating magnetic
field no more than 100 mT. The established average
ChRM directions are given in Table 1 (In situ and
Tilt corrected). The primary nature of the ChRM can
be substantiated by positive results of the reversals
and fold tests. The upper five of the studied sample
groups have a normal polarity, the mean direction
in stratigraphic coordinates: D = 319.2, I = 13.7,
269
Fig. 3. Typical orthogonal plots in tilt-corrected coordinates and corresponding NRM vs. T(AF) graphs based on the
results of T- and AF-demagnetization: (a) dacite from site 02ta-4 (sample number 02ta032); (b) rhyolite from site 02ta5 (sample number 02ta047); (c) andesite from site 02ta-6 (sample number 02ta057); gabbro-dolerite from site 02ta-9
(sample number 02ta098). Solid dots represent projections of vector endpoint on the horizontal plane, and the open ones
– on the vertical plane, the dashed line shows the stable ChRM component.
Table 1. Paleomagnetic directions and coordinates of virtual geomagnetic poles of the studied 960-Ma volcanogenic
formation from the Three Sisters Lake region
Table 1. Paleomagnetic directions and coordinates of virtual geomagnetic poles of the studied 960-Ma volcanogenic formation
from the Three Sisters Lake region
Site numbers,
rock type
02ta-3, gabbro-dolerite
02ta-4, dacite
02ta-5, rhyolite
02ta-6, andesite
02ta-7, gabbro-dolerite
02ta-9, gabbro-dolerite
02ta-10, gabbro-dolerite
Mean
n/N
10/11
8/10
10/10
7/10
9/12
8/10
10/10
In situ
D (°)
I (°)
Tilt corrected
D (°)
I (°)
70.0
152.5
327.5
209.4
300.4
85.8
94.6
264.2
323.9
319.5
320.1
310.7
321.7
138.9
144.8
16.3
16.9
14.4
18.1
2.3
-17.9
-17.6
320.0
14.8
86.1
88.0
89.4
80.5
-82.0
-69.1
-74.0
81.2
k
α95
374.1
118.9
75.4
92.7
38.5
89.1
72.8
2.9
127.3
2.5
5.1
5.6
6.3
8.4
5.9
5.7
43.0
5.4
PLat
19.1
18.8
17.6
17.9
11.7
19.2
20.0
VGPole
PLong
322.7
327.3
326.5
336.5
324.1
328.2
322.0
dp/dm
1.3/2.6
2.7/5.3
2.9/5.7
3.4/6.5
4.2/8.4
3.2/6.1
3.1/5.9
8.3±1.9
8.6±3.8
7.3±4.1
9.3±4.7
1.2±5.9
9.2±4.4
9.0±4.3
17.8
326.8
A95=4.0
7.5±4.0
PL
Note: n/N – ratio of the number of samples, used in the statistics, to the total number of studied samples; D – declination in
degrees; I – inclination in degrees; k – precision parameter, α95 – 95% confidence limit, VGPole – the virtual geomagnetic pole
coordinates (the inverted positions of the poles are given): PLat – latitude, PLong – longitude, dp/dm – semiaxes of the
confidence circle of paleomagnetic pole; the mean pole is calculated as the average from the VGPole batch where A95 - 95%
confidence limit; PL – paleolatitude for the reconstructed block in northern hemisphere.
270
V. A. Vernikovsky
Fig. 4. The relative positions of the calculated Central Taimyr 960 Ma paleomagnetic poles (Table 1, circles) and APWP
of Siberia for the time period 1,045–950 Ma by (Pavlov et al., 2002) and a paleogeographic reconstruction (inset) of the
Central Taimyr margin and Siberia at 960 Ma.
k = 100.8, α95 = 7.7. A reverse polarity has been
established for two outcrops (sample numbers 02ta9 and 02ta-10, table 1) of gabbro-dolerite sill, with
the mean direction being: D = 141.9, I = –17.8, k =
414.8, α95 = 12.3. The angle between the means of
the normal and reverse polarity is γ = 4.8° with γс =
9.4° as the critical value. The precision parameter (k)
is significantly higher in the stratigraphic coordinates
(the ratio ks/kg = 43.8 is higher than the critical value
– 4.16 for n = 7 at the 99% confidence level), the
optimal concentration of magnetic directions (when
k is maximum) is found at 109 ± 4% untilting. The
correlation test (McFadden, 1990) is positive: the test
parameter (distribution function) ξ1 in stratigraphic
coordinates – 3.166 exceeds the critical value at
95% confidence level – 3.086, at the same time in
geographic (in situ) coordinates ξ1 is 2.552, which
is lower than the critical value. The main stage of
deformations of the island arc complexes of Central
Taimyr corresponds to the ~600 Ma boundary
(Vernikovsky, Vernikovskaya, 2001), therefore
we can safely assume that the 960 Ma age of the
established ChRM is pre-Ediacaran (630-542 Ma;
Walker and Geissman, 2009). In all probability the
ICAM VI Proceedings
ChRM corresponds to the time of formation of these
rocks at 960 Ma or in the Early Neoproterozoic.
The mean paleomagnetic pole for the Central
Taimyr rocks (Table 1; PLat=17.8, Plong=326.8,
A95=4.0) is close to synchronous poles for Uya
series sedimentary rocks in the Uchuro-Maya
region in the south-east of the Siberian craton,
hosting basic intrusions (Pavlov et al., 2002). The
age of those intrusions is substantiated by results
of Sm–Nd, 942±19 Ma (Pavlov et al., 2002) and
U–Pb, 947±7, 1005±4 Ma (Rainbird et al, 1998)
dating. The angular divergence in the poles position,
considering the confidence interval, is equal to
31.7°±4.3° in paleolongitude and – 8.7°±3.7° in
paleolatitude (Fig. 4). Consequently, the island arc
whose relicts are preserved in the modern structure
of the Three Sisters Lake region was located in some
distance away from the Taimyr margin of Siberia at
the time of its formation. Judging by the difference
in synchronous paleomagnetic latitudes, the Central
Taimyr island arc could have been separated from
the Siberian continent by a back-arc basin 550–
1,380 km wide (Fig. 4). During the back-arc basin’s
closure the arc must have been rotated for ~30°
271
clockwise. From these observations, we reach the
following conclusions:
1. 960 m.y. ago the paleo-island arc of Central
Taimyr was located in the subequatorial zone,
near the northern margin of Siberia, and had
a sublatitudinal strike. The sizes of the backarc basin that existed between the arc and
the continent at this time could reach 556–
1,378 km incorporating the estimated error in
paleomagnetic determinations.
2. The established age of the island arc in the Central
Taimyr indicates that the transformation of the
passive continental margin regime into an active
one in the north of Siberia took place as early as
the beginning of the Neoproterozoic (1 Ga). It
is of fundamental significance for paleotectonic
reconstructions of Siberia’s position within
the framework of Rodinia because it does not
allow the joining of the Taimyr margin with the
Canadian margin of Laurentia as it is assumed
in alternative reconstructions (Dalziel, 1991;
Hoffman, 1991).
3. The accretion of the island arc to the craton
incorporated mutual rotation around a vertical
axis, which implies the existence of a significant
strike-slip component in the kinematics of the
accretion process in the north of Siberia in the
Late Precambrian.
THE KARA MICROCONTINENT DURING
THE PALEOZOIC
The Kara microcontinent (or Kara plate) is
one of the largest fragments of the ancient Arctida
paleocontinent (Fig.1). Therefore the problems
related to the reconstruction of its formation,
kinematics and interactions with other continental
blocks are very important for the understanding of
the entire Arctic region. The Kara microcontinent’s
Precambrian basement is heterogeneous, which is
consistent with refraction velocities ranging from
5.7 to 7.1 km/s (Bogolepov et al., 1991). According
to individual seismic soundings, the crust thickness
in the Kara microcontinent may exceed 40 km, with
a 14–16-km-thick lower crust. The sedimentary
Fig. 5. (a) Paleomagnetic site mean directions in situ; and (b) tilt corrected; and (c) APWP for the Kara microcontinent
and its comparison with APWP for Siberia by (Pechersky and Didenko, 1995) and APWP for Baltica by (Torsvik et al.,
1996). Modified after (Metelkin et al., 2005).
272
V. A. Vernikovsky
section includes two units. The lower unit is up to
14 km thick and is apparently composed of Late
Cambrian–Ordovician and Silurian to Permian
carbonate, evaporate, and terrigenous deposits
(Kaban’kov, Sobolevskaya., 1981; Kaban’kov et al.,
1982). The upper unit, as thin as 2 km, consists of
Triassic–Jurassic sequences.
Analysis of geostructural, paleomagnetic,
geochronological and biostratigraphic data showed
that the Kara microcontinent was tectonically
isolated from neighboring continents in the Early
Paleozoic (Fig.5) and collided with Siberia at ~300
Ma or in the Carboniferous (Vernikovsky et al.,
1995; 1998; 2004; Metelkin et al., 2000; 2005; 2012,
Shipilov, Vernikovsky, 2010). The 300 Ma collision
is thought to have closed an oceanic basin that once
separated Kara from Siberia and the Central Taimyr
island arc that collided with the Siberian continent at
~600 Ma. The Northern Taimyr forms the collision
belt between the Kara microcontinent on the north
and the Central Taimyr/Siberian amalgamation on
the south. The absence of Middle-Late Paleozoic
ophiolite and island arc complexes in the Northern
Taimyr is therefore curious.
We propose a reconstruction of the Early
Paleozoic history of the Kara microcontinent as part
of the amalgamation of the Arctida paleocontinent.
Our reconstruction describes the mechanism of
the 300 Ma collision of Kara with Siberia and
the subsequent collision-caused deformation
processes in the amalgamation of the greater Pangea
supercontinent that was culminated in Permian time
or ~280 Ma (Fig. 9) (Metelkin et al., 2011; 2012,
Vernikovsky et al., 2011).
Our paleotectonic analysis is based on
paleomagnetic and geochronological data. The
results indicate that the collision between Siberia
and the Kara microcontinent was an oblique event.
The orogen that was formed can be characterized as a
transform orogen (Metelkin et al., 2005). During the
Early Paleozoic and prior to the collision Kara moved
northward on a system of large transform faults from
the sub-tropic zone of the southern hemisphere to the
subequatorial latitudes of the northern hemisphere
while at the same time rotating counter clockwise,
whereas the Siberian plate underwent a clockwise
rotation (e.g., Fig. 8, 450 Ma).
The oppositely directed rotation of the interacting
ICAM VI Proceedings
Kara and Siberian continental masses led to their
oblique convergence and “soft” collision. In the Late
Silurian – Devonian, when there still was a “lens” of
oceanic crust between the Siberian and Kara plates
(Fig. 6, 430–400 Ma; Fig. 8, 420 Ma), the margins
of the converging continents escaped significant
shortening while sliding along the transforms and
maintaining intact margins (Fig. 6, 430–400 Ma).
It is possible that the oceanic crust was partially
subducted beneath the Siberian plate; however the
strike-slip processes were dominating. As a result the
supra-subductional geologic complexes are lacking.
The continent-microcontinent collision took place
in the Late Carboniferous and culminated in the
Permian (Fig. 6, panel for Carboniferous-Permian).
The Carboniferous-Permian (300–260 Ma) collision
re-deformed the Central Taimyr island-arc complex
that was originally deformed when it accreted to the
Siberian continent at 600 Ma.
In the course of oblique collision there was a
thickening of the crust, accompanied by folding which
migrated to the south-west (in modern geographic
coordinates), regional metamorphism, and the
formation of collisional granites (Vernikovsky et al.,
1995; Pease, 2001). As a result of compression in the
frontal part of the Kara tectonic domain there was a
gradual exhumation of the deeper parts of the crust
of the deformed plate.
The transform faults which controlled the
collision of the Kara and Siberia continental masses
gradually evolved into thrust faults as shortening
progressed. The oblique collision may have evolved
into a more orthogonal-directed collision because
of far-field interactions with the nearby continental
masses of Laurentia, Baltica, Alaska-Chukotka,
and Svalbard (Fig. 9). The most important among
these thrusts is the Main Taimyr fault zone (Fig.
6), which can be regarded as the main suture of
the Late Paleozoic Taimyr orogen. The progressing
compression and the associated crustal thickening
led to the “collapse” – fast thrusting and imbrication
of the crust and post-collisional granitoid magmatism
(Vernikovsky et al., 1998) on the Main Taimyr thrust
which separates the Central and North Taimyr zones.
This geodynamic paleoreconstruction for the
Kara microcontinent shows the need and significance
of such studies for the entire Arctic.
273
Fig. 6. A model for the structure transformation of the Siberian Taimyr margin during the interaction with the Kara
microcontinent. 1 – oceanic complexes; 2 – Early Precambrian complexes of the Kara microcontinent and Siberian
craton crystalline basement; 3–5 – Late Precambrian complexes of the Central Taimyr accretion zone: 3 – gneissic from
cratonic terranes, 4 – volcanogenic-sedimentary from island arc terranes with ophiolites, 5 – carbonate shelf of passive
continental margin terranes; 6 – Neoproterozoic-Cambrian flyschoid deposits of the Kara and Siberian continental
margins; 7 – Paleozoic mainly carbonate shelf deposits (Ordovician-Silurian on the Kara microcontinent and OrdovicianEarly Carboniferous on the Taimyr Siberian margin); 8 – Ediacaran-Early Carboniferous hemipelagic argilaceouscarbonate and black-schists deposits of the Pyasina-Faddey abyssal trough; 9 – Late Paleozoic mainly terrigeneous
deposits (Devonian and Carboniferous-Permian for Kara and Late Carboniferous-Permian for Southern Taimyr); 10
– Late Paleozoic (300–260 Ma) collisional granitoids; 11 – Triassic sandy-agrillaceous deposits, including the trap
complex in the front of the Late Paleozoic orogen (basal horizons of the Mesozoic-Cenozoic Yenisey-Khatanga basin).
274
V. A. Vernikovsky
SPECULATIVE PALEOTECTONIC RECONSTRUCTIONS FOR THE ARCTIDA PALEOCONTINENT AND THE GREATER ARCTIC
DURING LATE NEOPROTEROZOIC –
PERMIAN TIME
Cryogenian (~750 Ma)
The Cryogenian is marked by the breakup stage
of Rodinia – the supercontinent that formed around 1
Ga. According to current interpretations, the breakup
of Rodinia began as soon as ~950 Ma and continued
for a very long time until the Ediacaran (630–542
Ma) (Li et al., 2008). We join with Li et al., (2008)
in believing that most of the classic Arctida blocks
were composited into a continuous belt from
fragments originating in diverse settings including
the present-day northern margin of Laurentia, the
former (750–650 Ma) southern margin of Siberia,
and the present-day north-eastern margin of Baltica
(Fig. 7, 750 Ma).
The Svalbard plate has a Grenvillian (1.3–
1.0 Ga) basement, which has been confirmed by
the identification of Grenvillian complexes on
Spitsbergen (Gee et al., 1995) and on Novaya
Zemlya (Korago et al., 2004). This allows us to
infer the formation of Svalbard from collisional
events during the establishment of Rodinia. On
the basis of paleomagnetic data, Baltica is usually
positioned in paleoreconstructions in such a way that
in modern geographic coordinates the Grenvillian
Sveconorwegian structures serve as the northern
“ending” of the Grenvillian structures of Laurentia’s
eastern margin. The Meso-Neoproterozoic fold
belts of Amazonia are oriented in a linear fashion
along Laurentia’s Grenvillian margin (Cawood and
Pisarevsky, 2006). In this context it is logical to
suppose that the Svalbard orogen structures form the
northern (present-day) extension of the Grenville belt
that marks the collisions between Laurentia, Baltica,
and Amazonia that formed northern Rodinia.
Paleoproterozoic(?) crystalline complexes of
the Kara microcontinent basement are known on
the Severnaya Zemlya archipelago (Proskurnin,
1999; Proskurnin and Shul’ga, 2000) and in the
northern part of the Taimyr Peninsula (Vernikovsky
and Vernikovskaya, 2001). The sedimentary
cover on the Kara microcontinent is floored by
Late Neoproterozoic flyschoid deposits which
are overlain by a Paleozoic (Ordovician to Early
ICAM VI Proceedings
Carboniferous) sequence composed of carbonates,
evaporates and terrigenous formations that indicate
an epicontinental shelf regime. The structure of the
gravity, magnetic, and other geophysical fields for
the Kara microcontinent differ significantly from
adjacent plates or blocks. The Kara microcontinent
thus appears to form an independent block with a
distinct internal structure.
Despite the distinctive differences between
the geologic and geophysical structures of the
Kara microcontinent and the Svalbard plate,
the emplacement history and evolution of their
modern margin (St. Anna Trough and North
Siberian Sill) display a characteristic dextral strikeslip component, which was probably inherited
from the Neoproterozoic-Paleozoic transform
boundary between Svalbard and Kara (Shipilov
and Vernikovsky, 2010). From these observations
we speculate that in the Meso-Neoproterozoic
(Cryogenian) structure of Arctida (during the
formation of Rodinia) the Kara microcontinent was
located between the Greenland-Ellesmere block and
the Svalbard block, from the latter possibly separated
by a strike-slip fault system (Fig. 7, 750 Ma).
In our reconstruction the Early Precambrian
structures of Arctida’s Greenland-Ellesmere block
correspond to their current position near the Canadian
margin of Laurentia. Our reconstruction infers that
the Alaska-Chukotka block was located in close
proximity to the Greenland-Ellesmere block and
they did not change their positions as the northern
(present-day) margin of Laurentia throughout the
period 750–255 Ma. The detachment of the AlaskaChukotka tectonic element from Laurentia occurred
in the Jurassic (202–146 Ma), as part of the opening
of the Canada basin (Grantz et al., 1998, Lawyer et
al., 2002, Alvey et al., 2008). The Alaska-Chukotka
tectonic element later collided with the VerkhoyanKolyma Siberian plate along the South Anyui
(Novosibirsk-Chukotka) suture (Sokolov et al.,
2002; 2009).
Unlike earlier models, our reconstruction
does not include the New Siberian Islands and the
Laptev Sea continental shelf (New Siberian block)
in the structure of Meso-Neoproterozoic Arctida.
The Neoproterozoic-Paleozoic evolution of the New
Siberian block took place in a passive continental
margin setting (Kuzmichev, 2009). The Paleozoic
275
Fig. 7. Plate tectonic reconstructions for the evolution of Arctida and its dispersed fragments (Chukotka-Alaska, Kara,
and Svalbard) from the Neoproterozoic to the Early Ordovician. 1 – continental masses; 2 – continental blocks of
Arctida; 3 – oceanic basins; 4 – inferred position of spreading zones; 5 – active continental margins; 6 – general strike
of the transform/strike-slip zones with indicated strike-slip kinematics.
geological complexes that exist on the New Siberian
Islands are amazingly similar to the deposits of
the Cis-Verkhoyan and South-Taimyr margins of
Siberia. The lack of a pronounced tectonic suture in
the Laptev shelf allows us to infer the genetic unity
of the Paleozoic complexes of the New Siberian
block and the north-eastern Siberian margin. Thus in
our model in the Cryogenian the New Siberian block
276
was located far from the other Arctida blocks and
during the Paleozoic it evolved as a part of the northeastern (in geographic coordinates) Siberian margin.
According to our model, strike-slip
displacements took a major role in the process of
Rodinia’s breakup and basically conditioned the
tectonic dispersal of the supercontinent (Metelkin
et al., 2007, 2012; Vernikovsky et al., 2009). The
V. A. Vernikovsky
accepted position of the Siberian craton is based on
paleomagnetic data for the south of Siberia (Metelkin
et al., 2007, 2012) and is supplemented by data for
the Taimyr margin (Vernikovsky et al., 2011). The
latter indicate that Central Taimyr island arcs were
situated to the north from the Arctic margin of
Siberia since 960 Ma. Consequently, at the time of a
unified Rodinia (> 950 Ma) and later its breakup the
Siberian northern margin should have been facing a
paleo-ocean (Fig. 7, 750 Ma). The paleogeographic
position of the Arctida subcontinent was to the southwest relative to Siberia and straddling the equator
(Fig. 7, 750 Ma). The position of Kara, Svalbard and
Alaska–Chukotka within Arctida is debatable and
taken from (Li et al., 2008).
Cryogenian-Ediacaran (~650 Ma)
By the beginning of Ediacaran the Arctida
subcontinent on the northern margin of Laurentia
had moved south of the equator to subtropical
latitudes (Fig. 7, 650 Ma). Some Arctida blocks were
probably involved at this time in the breakup and
dispersal of Rodinia, including the detachment of
Baltica from Rodinia. Many examples show that the
Rodinia breakup was accompanied by the shredding
of the Rodinia margin into independent terranes such
as the Kara microcontinent and the Svalbard plate.
At the base of the Paleozoic sedimentary cover of
those plates Late Precambrian riftogenic troughs
and basins are present, which are clearly revealed by
seismic measurements (Shipilov and Vernikovsky,
2010).
At the same time on the eastern periphery of
Baltica (Timan-Ural margin) the evolution of an
active subduction zone can be inferred (Kuznetsov
et al., 2007). Oblique subduction on one side of the
Svalbard plate and extension on the other caused a
transform regime of its displacement and interaction
with the Kara plate.
Early Cambrian (~540 Ma)
Traces of the Cadomian orogenic event on the
territory of Barentsia (Puchkov, 2003; Kuznetsov
et al., 2007) in our opinion relate directly to the
evolution of the Arctida structures. We believe this
event to be a result of the collision between the Timan
margins of Baltica (present-day northeast margin)
with the Svalbard plate. From this time Barentsia was
ICAM VI Proceedings
joined to the East-European paleocontinent (Fig. 7,
540 Ma). The collision was structurally manifested
in the formation of the divergent Timan-Pechora
orogen. Its existence is confirmed by a deep cut-out
of the Late Precambrian complexes in the basement
of the Timan-Pechora sedimentary basin and by an
outstanding unconformity in the base of the Paleozoic
sedimentary cover (Kuznetsov et al., 2007). The 540
Ma collision was accompanied by the emplacement
of I-type granitoid plutons, characterized by isotopic
dates from 695 to 515 Ma (Kuznetsov et al., 2007).
Kara continued to experience a mainly transform
displacement relative to Svalbard. The transform/
strike-slip regime characterized the entire northeastern Siberian margin (in geographic coordinates)
and its displacements relative to distant Laurentia
and Baltica. On the boundary between Laurentia and
Baltica the Iapetus Ocean began to open (Fig. 7, 540
Ma).
Late Cambrian – Early Ordovician (~500 Ma)
By the Cambrian-Ordovician boundary (488
Ma) an active spreading regime widened the Iapetus
oceanic basin (Fig. 7, 500 Ma). The breakup of the
continental crust along the eastern (in present-day
coordinates) Baltica margin and the formation of
the Ural oceanic basin began at this time (Puchkov,
2003). Thus Baltica on almost all its periphery
(except the north) was surrounded by young oceanic
spreading centers whose growth dynamic set up a
counter-clockwise rotation of the plate, which is
confirmed by paleomagnetic data (Torsvik et al.,
1991; Cocks and Torsvik, 2002). The northern
Baltica margin including Svalbard was separated
from Siberia by large-scale strike-slip faults, which
caused a gradual drift of the Kara block towards
Siberia (Metelkin et al., 2005).
Late Ordovician (~450 Ma)
The Iapetus oceanic basin began to close at the
end of the Middle Ordovician. Active subduction
occurred widely on all of the margins of the
continents that surrounded the Iapetus Ocean.
Baltica began its movement across Iapetus toward
Laurentia. The Svalbard-Baltica margin and Kara
located on its periphery were drawn significantly
closer to the Taimyr margin of Siberia by mainly
multidirectional rotation of these continental masses.
277
Fig. 8. Plate tectonic reconstructions for the evolution of Arctida’s dispersed fragments from the Late Ordovician to the
Early Carboniferous. See legend keys on Fig. 7.
This entire system continued its general drift towards
the equator (Fig. 8, 450 Ma and 420 Ma).
Late Silurian – Late Devonian (~420–380 Ma)
During this time the collision between Laurentia
and Baltica (Laurussia) took place (Golonka et al.,
2003). Along with the formation of the Scandinavian
orogen the Caledonian orogeny also affected
Svalbard and the north-eastern Greenland margin,
later spreading along the Greenland-Ellesmere area
278
of Laurentia. Thus, by the end of the Silurian the
Ellesmere-Alaskan margin of Laurentia we infer the
existence of an active subduction zone where the
relicts of the Iapetus Ocean were consumed. The
Kara microcontinent already was approaching the
Taimyr margin of Siberia (Metelkin et al., 2005). The
early stages of the Kara-Siberia collision occurred
along a transform fault. The inferred transform fault
collision mechanism does not exclude the existence
of oceanic crust fragments between the Siberian
V. A. Vernikovsky
continent and the Kara microcontinent. Apparently
there also existed a narrow space of oceanic crust
between Svalbard and Kara. The Ural margin of
Baltica and the south-western Siberian margin were
characterized by intense subduction magmatism,
which indicates the closure of the Ural and PaleoAsian oceanic basins that were separated by the
Kazakhstan plate (Fig. 8, 420 Ma). The collisions
of the Siberian and Baltic plates took place along
strike-slip faults within their modern arctic margins.
As a result it was already the mid-Paleozoic (~380
Ma) when the component Arctic blocks of Arctida
were reassembled into their Cryogenian (~750
Ma) configuration. By this time the Arctida blocks
were located near the equator (Fig. 8, 380 Ma). By
the end of the Devonian the Arctida assemblage
formed a continental “bridge” between Siberia and
Laurussia (Laurentia/Baltica). According to our
reconstructions and available paleomagnetic data
for the Early Paleozoic of the Kara microcontinent
(Metelkin et al., 2005) we are inclined to believe that
the Siberian margin in the Silurian-Devonian did not
have any common boundaries with Laurussia. On
the west, an embayment of the Paleo-Pacific Ocean
separated the Siberian and Laurussia continental
blocks during subsequent Paleozoic evolution. The
Siberia and Laurussia plates were closest to each
other by the end of the Silurian. The transform
regime was dominating along all continental
margins of Arctida at ~380 Ma. Strike-slip faults
accommodated the sliding of Siberia and Kara to
the east along the north-western (in paleogeographic
coordinates) margin of Laurussia. This displacement
widened the embayment facing the Paleo-Pacific
Ocean into a wide marginal sea basin lapping the
margins of Alaska-Chukotka, Svalbard, Kara, and
New Siberia–Cis-Verkhoyan. It is probable that the
inferred strike-slip displacements were driven by
seafloor spreading Paleo-Pacific Ocean. To the east,
subduction and the closing of the Paleo-Pacific and
Ural Oceans added to the retreat of Siberia (and the
Arctic blocks sutured to its margin) away from the
Alaska-Chukotka margin of Laurentia.
Early Carboniferous (~355 and 330 Ma) and Late
Carboniferous (~305 Ma)
The Carboniferous period witnessed the closing
of the oceanic basins that divided the continental
ICAM VI Proceedings
masses of Laurussia (Baltica and Laurentia), Siberia,
and the Kazakhstan composite terrane. These
collisions culminated with the formation of Laurasia
– the supercontinent that along with Gondwana
formed Pangea at the Carboniferous-Permian
boundary (Fig. 9, 280 Ma) (Zonenshain et al., 1990;
Golonka, 2002).
At the beginning of the Carboniferous (355
Ma) the main blocks of Arctida (e.g., the AlaskaChukotka, Svalbard, Kara, and New Siberian blocks)
and the related continental margins of Laurentia,
Baltica and Siberia occupied the space between
the equator and 30° N. This entire paleo-shelf was
tectonically stable and underwent a slow “opening”
caused by the eastward retreat of Siberia. The main
cause for this retreat probably was seafloor spreading
in the Paleo-Pacific Ocean. The closing of the PaleoAsian and Ural Oceans as well as the progressive
narrowing of the Rheic and Paleo-Tethys Oceans was
essentially complete by ~305 Ma. These collisions
and the interactions with the Paleo-Pacific Ocean
on the western side of the Laurasian continental
agglomerate contributed to the transform-fault
regime of the paleo-shelf described above and to the
clockwise rotation of the system (Fig. 8, 355 Ma and
Fig. 9, 330 Ma).
By the Late Early Carboniferous (330 Ma) all
the continents continued drifting northwards, moving
closer to each other. for the final amalgamation of
continental masses into a unified supercontinent
began in Late Carboniferous time (Fig. 9, 305
Ma). Subduction of the Ural Ocean at the northeast
Baltica margin was completed (Puchkov, 2003). The
Paleo-Asian Ocean collapsed in a regime of oblique
subduction (Dobretsov, 2003; Windley et al., 2007).
At the end of the Early Carboniferous(Fig. 9, 330
Ma), collision tectonics began at the Taimyr margin
of Siberia (Vernikovsky et al., 1995; Vernikovsky,
1996). At Taimyr, the collision proceeded as a soft
interaction between sialic masses in oblique impact
conditions with them rotating relatively to each other
(Metelkin et al., 2005, 2012). Geochronological data
indicates that as early as in the Late Carboniferous
(305 Ma) syn-collisional calk-alkaline granitoids
began to intrude Taimyr (Vernikovsky et al., 1995;
Pease, 2001). Paleomagnetic data, described above,
forms the chief evidence for the inferred strike-slip
component of Taimyr deformation. Thus, large-
279
Fig. 9. Plate tectonic reconstructions for the evolution of Arctida and its dispersed fragments from the Early Carboniferous
to the Late Permian. See legend keys on Fig. 7.
scale strike-slip fault zones, along which Kara “slid”
during the entire Paleozoic, in the end led to the
collision between the Kara microcontinent and the
Siberian continent and subsequent formation of the
fold-and-thrust structure of the Taimyr – Severnaya
Zemlya folded area (located in Fig. 1b).
Early to Late Permian (280–255 Ma)
At the beginning of the Late Carboniferous (Fig.
9, 305 Ma) the main continental collisions involved
in the formation of Pangea had already started
(Zonenshain et al., 1990; Golonka, 2002; Metcalfe,
280
2002; Dobretsov, 2003). The CarboniferousPermian boundary (280 Ma) is the time when the
Laurasia and Gondwana blocks united in a single
supercontinent – Pangea (Fig. 9, 280 Ma). The
deformations caused by the collision and orogenic
events continued within Laurasia, related mostly to
strike-slip displacements along old sutures. Available
paleomagnetic data indicate that the intraplate strikeslip displacements between rigid tectonic units of
Eurasia (the Siberian and East European cratons)
continued until the Cenozoic (Metelkin et al., 2010,
2012). By the end of the Permian (Fig. 9, 255 Ma)
V. A. Vernikovsky
the mainly transform-fault-driven amalgamation of
the Kara – New Siberian and Svalbard – Novaya
Zemlya continental margins into a single shelf
structure was accomplished (Shipilov, 2003; 2008).
The collisions caused the curved structure of the
Pay-Khoy – Novaya Zemlya area (Korago et al.,
1992; Scott et al., 2010).
Thus the Permian-Triassic boundary can be
considered as the time of the second formation
of Arctida or “Arctida-II”. Arctida-II is located
in Pangea’s northern edge near the 60th parallel,
occupying the moderate and sub-polar regions
of the Northern Hemisphere (Fig. 9, 255 Ma).
Subsequently, in the Mesozoic as a result of the
opening of the Amerasian basin, the large AlaskaChukotka block was rifted away from the GreenlandEllesmere margin (Grantz et al., 1998). Its collision
with the Cis-Verkhoyan Siberian margin in the
Cretaceous along the South Anyui suture (Drachev
et al., 1998; Sokolov et al., 2002; 2009; Kuzmichev,
2009) established the main structural features of the
current arctic shelves of the Eurasian and NorthAmerican continents (Natalin, 1999; Khain et al.,
2009).
CONCLUSION
We infer the existence at two different times of
two Arctic subcontinents comprised of essentially
the same crustal fragments. The first subcontinent,
“Arctida-I” broke apart and the fragments were
dispersed through independent plate movement
paths before being reassembled as the second
subcontinent, “Arctida-II.”
Arctida-I
was
an
amalgamation
of
Mesoproterozoic terranes that “welded” together
elements of Laurentia, Siberia and Baltica within
the Rodinia supercontinent at 1 Ga. The Rodinia
disintegration caused the breakup of Arctida-I into
independent tectonic fragments which experienced
highly diverse displacement paths over the
next 720 million years (1,000 to 280 Ma). The
evolution of Neoproterozoic and Paleozoic oceanic
basins between these tectonic fragments led to
their reorganization into a new configuration in
Arctida-II – a Late Paleozoic subcontinent which
again “welded” together the continental masses
of Laurentia, Siberia and Baltica within Pangea.
The breakup of Arctida-II in the Mesozoic and the
ICAM VI Proceedings
Cenozoic with the formation of the north Atlantic
basin and the Amerasian and Eurasian basins of the
Arctic Ocean led to a significant redistribution of the
continental masses, especially in the north-eastern
part of the modern Arctic and to the formation of the
modern shelves of the Eurasian and North-American
continents.
Our paleotectonic reconstructions will of course
be improved after further investigations. For this
purpose complex geostructural, geochronological,
paleontological and especially paleomagnetic data
will be of paramount importance.
REFERENCES
Alvey, A., Gaina, C., Kusznir, N.J., Torsvik, T.H.,
2008. Integrated crustal thickness mapping and
plate reconstructions for the high Arctic. Earth
and Planetary Science Letters, 274, 310–321.
Butler, R.F., 1992. Paleomagnetism: magnetic
domains to geologic terrains. Backwell Sci.
Publ., Oxford, 319.
Bezzubtsev, V.V., Zalyaleev, R.Sh., Sakovich, A.B.,
1986. Geological map of Gorny Taimyr Scale 1:
500,000, Explanatory Note. Krasnoyarsk, 177
(in Russian).
Bogolepov, A.K., Shipilov, E.V., Yunov, A.Y., 1991.
Geophysical characteristics of the Barents-Kara
region deep structure. Proc. Intern. Seminar on
Tectonics of the Barents Sea – Kara Sea Region.
NIIMorgeophysica. IKU, Inter. Groupe,
Norvey, Trondheim.
Cawood, P.A., Pisarevsky, S.A., 2006. Was
Baltica right-way-up or upside-down in the
Neoproterozoic? J. Geol. Soc. London, 163,
753–759.
Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography
from 500 to 400 million years ago: a faunal and
palaeomagnetic review. J. Geol. Soc., 159, 631–
644.
Cocks L.R.M., Torsvik T.H. Siberia, the wandering
northern terrane, and its changing geography
through the Palaeozoic // Earth Sci. Rev., 2007,
v.82, p. 29–74.
Condie, K.C., Rosen,O.M., 1994. Laurentia–Siberia
connection revisited. Geology, 22, 168–170.
Dalziel, I.W.D., 1991. Pacific margins of Laurentia
and East Antarctica–Australia as a conjugate
281
rift pair: evidence and implications for an
Eocambrian supercontinent. Geology, 19, 598–
601.
Dalziel, I.W.D., 1997. Neoproterozoic – Paleozoic
geography and tectonics: review, hypothesis
and environmental speculation. Geol. Soc. Am.
Bull., 109, 16–42.
Drachev, S.S., Savostin, L.A., Groshev, V.G.,
Bruni, I.E., 1998. Structure and geology of the
continental shelf of the Laptev Sea, Eastern
Russian Arctic. Tectonophysics, 298, 357–393.
Gee, D.G., Johansson, A., Ohta, Y., Tebenkov, A.M.,
Krasilschikov, A.A., Balashov, Y.A., Larionov,
A.N., Gannibal, L.F., Ryungenen, G.I., 1995.
Grenvillian basementand a major unconformity
within the caledonides of nordaustlandet,
Svalbard. Prec. Research, 70(3–4), 215–234.
Golonka, J., 2002. Plate tectonics map of the
Phanerozoic. In: Kiesling, W., Flügel, E.,
Golonka, J. (eds.), Phanerozic reef patterns.
SEPM Spec. Publ., 72, 21–75.
Golonka, J., Bocharova, N.Y., Ford, F., Edrich,
M.E., Bednarczyk, J., Wildharber, J., 2003.
Paleogeographic reconstruction and basins
development of the Arctic. Mar. Petrol. Geol.,
20(3–4), 211–248.
Golonka J., Krobiski M., Pajak J., Nguyen Van
Giang, Zuchiewicz W. Global Plate Tectonics
and Paleogeography of Southeast Asia. Faculty
of Geology, Geophysics and Environmental
Protection, AGH University of Science and
Technology, Arkadia, Krakow, Poland, 2006,
128 p.
Grantz, A., Clark, D.L., Phillips, R.L., Srivastava,
S.P., Blome, C.D., Gray, L.B., Haga, H., Mamet,
B.L., McIntyre, D.J., McNeil, D.H., Mickey,
M.B., Mullen, M.W., Murchey, B.I., Ross,
C.A., Stevens, C.H., Silberling, N.J., Wall, J.H.,
Willard, D.A., 1998. Phanerozoic stratigraphy
of Northwind Ridge, magnetic anomalies in
the Canada Basin, and the geometry and timing
of rifting in the Amerasia basin, Arctic Ocean.
Geol. Soc. Am. Bull., 110, 801–820.
Dobretsov, N.L., 2003. Evolution of structures of the
Urals, Kazakhstan, Tien Shan, and Altai-Sayan
region within the Ural-Mongolian Fold Belt
(Paleo-Asian ocean). Russ. Geol. Geophys.,
44(1–2), 5–27.
282
Enkin, R.J., 2003. The direction-correction tilt test:
an all-purpose tilt/fold test for paleomagnetic
studies. Earth and Planetary Science Letters,
212, 151–166.
Hoffman, P.F., 1991. Did the breakout of Laurentia
turn Gondwana inside out? Science, 252, 1409–
1412.
Kaban’kov, V.Ya., Sobolevskaya, R.F., 1981. Late
Precambrian-Early Paleozoic history of the
Taimyr–Severnaya Zemlya folded area. In:
Tectonics of the Baikalian (Riphean) Strata of
Siberia. IGiG SO AN SSSR, Novosibirsk, 55–
62 (in Russian).
Kaban’kov, V.Ya., Rogozov, Yu.G., Makar’ev,
A.A., 1982. Upper Proterozoic stratigraphy
of Bolshevik Island (Severnaya Zemlya). In:
Geology of the Severnaya Zemlya Archipelago.
Sevmorgeo, Leningrad, 5–22 (in Russian).
Khain, V.E., Gusev, G.S., Khain, E.V., Vernikovsky,
V.A.,Volobuyev, M.I., 1997. Circum-Siberian
Neoproterozoic Ophiolite Belt. Ofioliti, 22(2),
195–200.
Khain, V.E., Polyakova, I.D., Filatova, N.I., 2009.
Tectonics and petroleum potential of the East
Arctic Province. Russ. Geol. Geophys., 50(4),
334–345.
Kirschvink, J.L., 1980. The least-squares line and
plane and the analysis of paleomagnetic data.
Geophysical Journal of the Royal Astronomical
Society, 62, 699–718.
Korago, E.A., Kovaleva, G.N., Il’in, V.F., Pavlov,
L.G., 1992. The Tectonics and Metallogeny
of the Early Kimmerides of Novaya Zemlya.
Nedra, Leningrad.
Korago, E.A., Kovaleva, G.N., Lopatin, B.G., Orgo,
V.V., 2004. The Precambrian Rocks of Novaya
Zemlya. Geological Society, London, Memoirs,
30, 135–143.
Kurenkov S.A., didenko A.N., Simonov V.A.,
Lubnina N.V. Geodynamic formation conditions
of the paleospreading complexes of the UralMongolian belt / Proc. GIN. Iss. 561, Studies
on regional tectonics. V. 2: Kazakhstan, TianShan, Polar Ural, M.: Nauka, 2005. 249 P. (in
Russian)
Kuzmichev, A.B., 2009. Where does the South
Anyui suture go in the New Siberian Islands
and Laptev Sea?: Implications for the Amerasia
V. A. Vernikovsky
basin origin. Tectonophysics, 463, 86–108.
Kuznetsov, N.B., Soboleva, A.A., Udoratina, O.V.,
Gertseva, O.V., Andreichev, V.L., 2007. PreOrdovician tectonic evolution and volcano–
plutonic associations of the Timanides and
northern Pre-Uralides, northeast part of the East
European Craton. Gondwana Research, 12,
305–323.
Lawver, L.A., Grantz, A., Gahagan, L.M., 2002. Plate
kinematic evolution of the present Arctiregion
since the Ordovician. In: Miller, E.L., Grantz,
A., Klemperer, S.L. (eds.) Tectonic Evolutionof
the Bering Shelf–Chukchi Sea–Arctic Margin
and Adjacent Landmasses. Geol. Soc. Amer.,
Spec. Pap., 360, 333–358.
Lawver L.A., Gahagan L.M. and Norton I.
Palaeogeographic and tectonic evolution of
the Arctic region during the Palaeozoic //
Geological Society, London, Memoirs, 2011;
v.35; p. 61-77.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson,
A., Waele, B.De., Ernst, R.E., Fitzsimons
I.C.W., Fuck, R.A., Gladkochub, D.P., Jacobs, J.,
Karlstrom, K.E., Lul, S., Natapov, L.M., Pease,
V., Pisarevsky, S.A., Thrane, K., Vernikovsky,
V., 2008. Assembly, configuration, and breakup history of Rodinia: A synthesis. Precam.
Res., 160, 179–210.
Meert J.G., Torsvik T.H. The making and unmaking
of a supercontinent: Rodinia revisited //
Tectonophysics, 2003, v.375, p.261–288.
McElhinny M.W., MacFadden P.L. Paleomagnetism:
continents and oceans. San Diego, Academic
Press, CA, 2000, 386 p.
McFadden, P.L., 1990. A new fold test for
paleomagnetic studies. Geophys. J. Int., 103,
163–169.
McFadden, P.L., McElhinny, M., 1990. Classification
of reversal test in paleomagnetism. Geophys. J.
Int., 103, 725–729.
Meert J.G., Torsvik T.H., 2003. The making
and unmaking of a supercontinent: Rodinia
revisited. Tectonophysics, 375, 261–288.
Metcalfe I., 2002. Permian tectonic framework and
palaeogeography of SE Asia. J. Asian Earth
Sci., 20, p.551–566.
Metelkin, D.V., Kazansky, A.Yu., Vernikovsky,
V.A., Gee, D.G., Torsvik, T., 2000. First
ICAM VI Proceedings
paleomagnetic data on the Early Paleozoic
rocks from the Severnaya Zemlya Archipelago
and their geodynamic interpretation. Russ.
Geol. Geophys., 41(12), 1767–1772.
Metelkin, D.V., Vernikovsky, V.A., Kazansky,
A.Yu., Bogolepova, O.K., Gubanov, A.P., 2005.
Paleozoic history of the Kara microcontinent
and its relation to Siberia and Baltica:
paleomagnetism, paleogeography and tectonics.
Tectonophysics, 398, 225–243.
Metelkin, D.V., Vernikovsky, V.A., Kazansky, A.Yu.,
2007. Neoproterozoic evolution of Rodinia:
constraints from new paleomagnetic data on
the western margin of the Siberian craton. Russ.
Geol. Geophys., 48(1), 32–45.
Metelkin, D.V., Vernikovsky, V.A., Kazansky, A.Yu.,
2011. Siberia – from Rodinia to Eurasia. In
Closson, D. (ed.) Tectonics, In Tech, Rijeka,
Croatia, 103–136.
Metelkin, D.V., Vernikovsky, V.A., Kazansky,
A.Yu., 2012 Tectonic evolution of the Siberian
paleocontinent from the Neoproterozoic to
the Late Mesozoic: paleomagnetic record and
reconstructions. Russ. Geol. Geophys., 53(7),
681–694.
Natalin, B.A., Amato, J.M., Toro, J., Wright, J.E.,
1999. Paleozoic rocks of northern Chukotka
Peninsula Russian Far East: implications for
the tectonics of the Arctic region. Tectonics, 18,
977–1003.
Pavlov, V.E., Gallet, Y., Petrov, P.Yu., Zhuravlev,
D.Z., Shatsillo, A.V., 2002. Uy series and late
Riphean sills of the Uchur-Maya area: isotopic
and palaeomagnetic data and the problem of
the Rodinia supercontinent. Geotectonics, 36,
278–292.
Pease, V., 2001. East European Craton margin
source for the allochthonous Northern Terrane
of Taimyr, Arctic Siberia. EOS Transactions,
American Geophysical Union, 82, Fall Meeting
Suppl., Abstract T32B-0892.
Pease, V., Gee, D., Vernikovsky, V., Vernikovskaya,
A., Kireev, S., 2001. The Mamont-Shrenk
terrane: a Mesoproterozoic complex in the
Neoproterozoic accretionary belt of Central
Taimyr, Northern Siberia. Terra Nova, 13, 270–
280.
Pechersky D.M., Didenko A.N. The Paleoasian
283
ocean: petromagnetic and paleomagnetic
information concerning its lithosphere.
M.:OIFZ RAN, 1995. 298 P. (in Russian).
Pisarevsky, S.A., Natapov, L.M., Donskaya, T.V.,
Gladkochub, D.P., Vernikovsky, V.A., 2008.
Proterozoic Siberia: a promontory of Rodinia.
Precam. Res., 160, 66–76.
Powell, C.M., Li, Z.X., McElhinny, M.W.,
Meert, J.G., Park, J.K., 1993. Paleomagnetic
constraints on timing of the Neoproterozoic
breakup of Rodinia and the Cambrian formation
of Gondwana. Geology, 21, 889–892.
Proskurnin, V.F., 1999. Concerning the problem
of angular unconformities in the Upper
Precambrian and Lower Paleozoic of the
Severnaya Zemlya archipelago. In: Simonov,
O.N. (ed.). Bowels of Taimyr. Norilsk, 68–76,
(in Russian).
Proskurnin, V.F., Shul’ga, Yu.D., 2000. Geological
structure of the pre-Mesozoic basement,
magmatic formations. In: Gramberg, I.S.,
Ushakov, V.I. (eds.). Severnaya Zemlya – geology
and mineral resources. VNIIOkeangeologia, St.
Petersburg, 31–44 (in Russian).
Puchkov, V.N., 2003. Uralides and Timanides,
their structural relationship and position into
geological history of Ural-Mongolian foldthrust belt. Russ. Geol. Geophys., 44(112),
28–39.
Rainbird, R.H., Stern, R.A., Khudoley, A.K.,
Kropachev, A.P., Heaman, L.M., Sukhorukov,
V.I., 1998. U–Pb geochronology of Riphean
sandstone and gabbro from southeast Siberia and
its bearing on the Laurentia–Siberia connection.
Earth Planet. Sci. Lett., 164, 409–420.
Scotese, C.R., McKerrow, W.S., 1990. Revised world
map and introduction. In: McKerrow, W.S.,
Scotese, C.R. (eds.). Plaeozoic Paleogeography
and Biogeography, Geol. Soc. London Mem.,
12, 1–21.
Scotese C.R. Paleogeographic Atlas / PALEOMAP
Progress Report 90-0497, 1997, Department
of Geology, University of Texas at Arlington,
Arlington, TX, 45 pp.
Scott, R.A., Howard, J.P., Guo, L., Schekoldin, R.,
Pease, V., 2010. Offset and curvature of the
Novaya Zemlya fold-and-thrust belt, Arctic
Russia. In: Vining, B.A., Pickering, S.C. (eds.).
284
Petroleum Geology: From Mature Basins
to New Frontiers – Proceedings of the 7th
Petroleum Geology Conference, 645–657.
Shipilov, E.V., 2003. Interference of tectonicgeodynamic regimes in Permian- Triassic
evolution of the Arctic periphery of North
Eurasia. Dokl. Earth Sci. 393A(9), 1215–1219.
Shipilov, E.V., 2008. Generations of spreading
basins and stages of breakdown of Wegener’s
Pangea in the geodynamic evolution of the
Arctic Ocean. Geotectonics, 42(2), 105–124.
Shipilov, E.V., Vernikovsky, V.A., 2010. The
Svalbard–Kara junction: structure and
geodynamic history Russ. Geol. Geophys., 51,
1–13.
Smethurst, M.A., Khramov, A.N., Torsvik, T.H.,
1998. The Neoproterozoic and Paleozoic
paleomagnetic data for the Siberian Platform:
from Rodinia to Pangea. Earth Sci. Rev., 43, 1–
24.
Sokolov, S.D., Bondarenko, G.Ye., Morozov, O.L.,
Shekhovtsov, V.A., Glotov, S.P., Ganelin, A.V.,
Kravchenko-Berezhnoy, I.R., 2002. The South
Anyui suture: facts and problems to solve
In: Miller, E.L., Klemperer, S., Grantz, A.
(eds.). Tectonic evolution of the Bering shelf
– Chukchi sea – Arctic margin and adjacent
landmasses, Geol. Soc. of America Spec. Paper
360. 209–224.
Sokolov, S.D., Bondarenko, G.Ye., Layer, P.W.,
Kravchenko-Berezhnoy, I.R., 2009. South Anyui
suture: tectono-stratigraphy, deformations,
and principal tectonic events. Stephan Mueller
Special Publication Series “Geology, geophysics
and tectonics of Northeastern Russia: a tribute
to Leonid Parfenov”, 4, 201–221.
Torsvik T.H., Ryan P.D., Trench A., Harper D.A.T.,
1991. Cambrian-Ordovician paleogeography of
Baltica. Geology, 19, 7–10.
Torsvik, T.H., Smethurst, M.A., Meert, J.G., Van der
Voo, R., McKerrow, W.S., Brasier, M.D., Sturt,
B.A., Walderhaug, H.J., 1996. Continental
break-up and collision in the Neoproterozoic
and Palaeozoic – a tale of Baltica and Laurentia.
Earth Sci. Rev., 40(3–4), 229–258.
Torsvik T.H.and Van der Voo R. 2002, Refining
Gondwana and Pangea palaeogeography:
estimates of Phanerozoic non-dipole (octupole)
V. A. Vernikovsky
fields // Geophys. J. Int., v.151(3), p.771–794.
Uflyand, A.K., Natapov, L.M., Lopatin, V.M.,
Chernov, D.V., 1991. To the Taimyr tectonic
nature. Geotectonika, 6, 76–93, (in Russian).
Vernikovsky, V.A., 1996. Geodynamic evolution of
Taimyr fold area. Novosibirsk, Siberian Branch
RAS, SPC UIGGM, 203, (in Russian).
Vernikovsky, V.A., Vernikovskaya, A.E., Lyapunov,
S.M., Neimark, L.A., Proskurnin, V.F.,
Chernykh, A.I., 1994. Petrology, geochemistry,
and tectonic settings of plagiogranites of the
Chelyuskin Ophiolite Belt. Int. Geol. Rev., 36,
961–974.
Vernikovsky, V.A., Neimark, L.A., Ponomarchuk,
V.A., Vernikovskaya, A.E., Kireev, A.D.,
Kuzmin, D.S., 1995. Geochemistry and age of
collision granitoids and metamorphites of the
Kara microcontinent (Northern Taimyr). Russ.
Geol. Geophys., 36, 46–60.
Vernikovsky, V.A., Sal’nikova, E.B., Kotov, A.B.,
Ponomarchuk, V.A., Kovach, V.P., Travin, A.V.,
1998. Age of post-collisional granitoids, North
Taimyr: U-Pb, Sm-Nd, Rb-Sr and Ar-Ar data.
Dokl. RAN, 363, 375–378 (in Russian).
Vernikovsky, V.A., Vernikovskaya, A.E., 2001.
Central Taimyr accretionary belt (Arctic Asia):
Meso–Neoproterozoic tectonic evolution and
Rodinia breakup. Prec. Res., 110, 127–141.
Vernikovsky, V.A., Vernikovskaya, A.E., Pease,
V.L., Gee, D.G., 2004. Neoproterozoic orogeny
along the margins of Siberia. Geol. Soc. London
Mem., 30, 233–247.
Vernikovsky, V.A., Kazansky, A.Yu., Matushkin,
N.Yu., Metelkin, D.V., Sovetov, J.K., 2009. The
geodynamic evolution of the folded framing
and the western margin of the Siberian craton
in the Neoproterozoic: geological, structural,
sedimentological,
geochronological,
and
paleomagnetic data. Russ. Geol. Geophys., 50,
380–393.
Vernikovsky, V.A., Metelkin, D.V., Vernikovskaya,
A.E., Matushkin, N.Yu., Lobkovsky, L.I.,
Shipilov, E.V., 2011. Early stages of evolution
of the arctic margins (NeoproterozoicPaleozoic) and plate reconstructions. Origins
of Northeastern Russia: Paleomagnetism,
Geology, and Tectonics. ICAM-VI. Abstracts,
Fairbanks, Alaska, CD.
ICAM VI Proceedings
Vernikovsky, V.A., Metelkin, D.V., Vernikovskaya,
A.E., Sal’nikova, E.B., Kovach, V.P., Kotov,
A.B., 2011. The Oldest Island Arc Complex
of Taimyr: Concerning the Issue of the
Central Taimyr Accretionary Belt Formation
and Paleogeodynamic Reconstructions in the
Arctic. Doklady Earth Sciences, 2011, 436(2),
186–192.
Walker, J.D., and Geissman, J.W. (compilers), 2009,
Geologic time scale: Geological Society of
America, doi: 10.1130/2009.CTS004R2C.
Watson, G.S., and Enkin, R.J., 1993. The fold test
in paleomagnetism as a parameter estimation
problem. Geophys. Res. Lett., 20, 2135–2137.
Windley B.F., Alexeiev D., Xiao W., Krцner
A., Badarch G., 2007. Tectonic models for
accretion of the Central Asian Orogenic Belt. J.
Geol. Soc. Lond., 164(1), 31–47.
Wingate M.T.D., Giddings J.W. Age and
palaeomagnetism of the Mundine Well dyke
swarm, Western Australia implications for an
Australia-Laurentia connection at 755 Ma //
Precam. Res., 2000, v.100, p.335–357.
Zabiyaka, A.I., Zabiyaka, I.D., Vernikovsky,
V.A., Serdyuk, S.S., Zlobin, M.M., 1986.
Geological Structure and Tectonic Evolution
of Northeastern Taimyr. Nauka, Novosibirsk,
144p.
Zonenshain, L.P., Natapov, L.M., 1987. The
Arctic tectonic history. In: Topical Problems
of Geotectonics. Nauka, Moscow, 31–57 (in
Russian).
Zonenshain, L.P., Kuz’min, M.I., Natapov, L.M.,
1990. Geology of the USSR: A Plate Tectonic
Synthesis, Geodyn. Ser. 21, Page, B.M. (ed.),
AGU, Washington, D.C.
285